Heterostructured Bi2S3–Bi2O3 Nanosheets with a Built-In Electric

Feb 6, 2018 - In this work, heterostructured Bi2S3–Bi2O3 nanosheets (BS–BO) have been prepared through an easy water-bath approach. The formation ...
0 downloads 6 Views 2MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Article

Heterostructured Bi2S3-Bi2O3 Nanosheets with Built-In Electric Field for Improved Sodium Storage Wen Luo, Feng Li, Qidong Li, Xuanpeng Wang, Wei Yang, Liang Zhou, and Liqiang Mai ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.8b01613 • Publication Date (Web): 06 Feb 2018 Downloaded from http://pubs.acs.org on February 7, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Heterostructured Bi2S3-Bi2O3 Nanosheets with Built-In Electric Field for Improved Sodium Storage Wen Luo,a,b† Feng Li,a† Qidong Li,a Xuanpeng Wang,a Wei Yang,a Liang Zhoua and Liqiang Maia,c*

a

State Key Laboratory of Advanced Technology for Materials Synthesis and

Processing, Wuhan University of Technology, Wuhan 430070, China b

Laboratory of Chemistry and Physics: Multiscale Approach to Complex

Environments (LCP-A2MC), Institute of Jean Barriol, University of Lorraine, Metz 57070, France c

Department of Chemistry, University of California, Berkeley, California 94720,

United States



These authors contributed equally to this work.

*Corresponding author: [email protected]

1 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 25

Abstract: Constructing novel heterostructures has great potential in tuning the physical/chemical properties of functional materials for electronics, catalysis, as well as energy conversion and storage. In this work, heterostructured Bi2S3-Bi2O3 nanosheets (BS-BO) have been prepared through an easy water-bath approach. The formation of such unique BS-BO heterostructures was achieved through a controllable thioacetamide-directed surfactant-assisted reaction process. Bi2O3 sheets and Bi2S3 sheets, can be also prepared through simply modifying the synthetic recipe. When employed as the sodium-ion battery anode material, the resultant BS-BO displays a reversible capacity of ~630 mA h g-1 at 100 mA g-1. In addition, the BS-BO demonstrates improved rate capability and enhanced cycle stability compared to its Bi2O3 sheets and Bi2S3 sheets counterparts. The improved electrochemical performance can be ascribed to the built-in electric field in the BS-BO heterostructure, which effectively facilitates the charge transport. This work would shed light on the construction of novel heterostructures for high-performance sodium-ion batteries and other energy related devices. Keywords: Bi2S3-Bi2O3 nanosheets, heterostructure, built-in electric field, sodium-ion batteries, anode

2 ACS Paragon Plus Environment

Page 3 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

INTRODUCTION The demand for low-cost, high-efficiency energy storage technologies is greatly increasing to utilize renewable energy resources and reduce environmental pollution. Rechargeable lithium-ion batteries (LIBs) are undoubtedly one of the most successful candidates among various energy storage devices.1-3 Recently, sodium-ion batteries (SIBs) are emerging as promising alternative to commercial LIBs due to the abundant, inexpensive sodium resources and its similar redox chemistry to LIBs.4 Nevertheless, because the ionic radius of Na+ ions is much larger (102 pm) than that of Li+ ions (76 pm), most investigated electrode materials suffer from huge volume expansion during sodiation/de-sodiation processes, resulting in low specific capacity and limited cycle life.5-8 Hence, the pursuit of SIB electrode materials with high specific capacity and stable cycling performance remains a major challenge. Metal chalcogenides, such as Fe3S4,9 FeSe2,10 Sb2S311 and Sb2Se312 have attracted considerable attentions in sodium storage owing to their high specific capacity and intriguing layered structure. Bi2S3 is a well-known layered p-type semiconductor with a band gap of approximate 1.3 eV, and it displays enormous potentials in thermoelectrics,13, 14 photocatalysis,15-18 photodetectors,19-21 as well as electrochemical energy conversion and storage.22-24 For sodium storage, Bi2S3 is capable to display gravimetric and volumetric capacities of 625 mA h g-1 and 4250 mA h cm-3, respectively.25, 26 Bi2O3 is also a typical layered semiconductor with a band gap of 2.8 eV (n-type semiconductor).27 For sodium storage, the Bi2O3 possesses a high theoretical capacity of 690 mA h g-1.28, 29 Despite their high theoretical capacities, the 3 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 25

sodium storage performances of Bi2O3 and Bi2S3 are usually unsatisfactory. For instance, the Bi2O3/C composite displayed a short life span of only 20 cycles.29 The Bi2O3@rGO nanocomposite exhibited a capacity retention of 70.2% after 200 cycles.30 The binder-free and flexible Bi2O3/C delivered a specific capacity of 430 mA h g-1 after 200 cycles.31 The Bi2S3@CNT nanocomposite delivered a capacity of only 84.8 mA h g-1 after 60 cycles at 60 mA g-1.32 Further enhancement of the sodium storage performance of Bi-based anodes still remain a great challenge. Constructing heterostructures composed of at least two components with different band gaps have great potential in tuning the physical/chemical properties of functional materials.33-36 Due to the unique interface effect, many advantages have been realized in heterostructures in various energy storage devices. Of particular note that Guo et al. proposed a conceptual model of SnS/SnO2 heterostructures with built-in electric field and thus boosted charge transfer capability.37 Using ultrathin Bi2MoO6 nanosheet as an example, the same group interpreted the electric-field effect based on density functional theory calculations.38 Recently, Yu et al. confirmed that unbalanced charge distribution would occur within crystals and induce built-in electric fields, leading to boosted lithium ion transfer dynamics.39 Herein, inspired by built-in electric field effect and enhanced charge transfer in heterostructured materials, we designed Bi2S3-Bi2O3 (BS-BO) heterostructures and explored them in SIBs application. Detailed characterizations demonstrated that the formation of such unique BS-BO heterostructures was achieved through a controllable thioacetamide-directed surfactant-assisted reaction process. Bi2O3 sheets and Bi2S3 sheets can also be 4 ACS Paragon Plus Environment

Page 5 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

prepared by simply modifying the synthetic recipe. The BS-BO, as an anode for SIBs, delivers a reversible capacity of 630 mA h g-1 at 100 mA g-1. Besides, the BS-BO also demonstrates better cycling stability and rate capability compared with the Bi2O3 sheets and Bi2S3 sheets counterparts. The improved sodium storage performance makes the BS-BO heterostructures a promising anode for SIBs and the facile synthetic approach also provides new insights into the construction of novel heterostructures. EXPERIMENTAL SECTION Synthetic procedures For the synthesis of BS-BO heterostructures, 0.243 g Bi(NO3)3·5H2O and 0.125 mL HNO3 (14 M) were first added into 5 mL deionised water, leading to the formation of a milky suspension. Then a sulfurization agent solution was prepared by dissolving 0.125 g of hexadecyl trimethyl ammonium bromide (CTAB) and 0.075 g of thioacetamide (TAA) in 60 mL deionised water through ultrasonication. The above milky suspension was added into the sulfurization solution dropwise with constant stirring. The mixture was incubated at 30 °C for 3 h. The final product was collected by centrifugation and thoroughly washed with deionised water and ethanol for three times. The synthesis of Bi2S3 sheets is similar to that of BS-BO heterostructure except a higher amount of TAA (0.375 g) was used. If no TAA was used in the synthesis, Bi2O3 sheets were obtained through the incubation of milky suspension at 30 °C for 3 h. Characterization Field-emission scanning electron microscopy (FESEM) images were collected 5 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 25

with a JEOL-7100F microscope. X-ray diffraction (XRD) patterns were obtained with a D8 Advance X-ray diffractometer, using Cu Kα radiation (λ = 1.5418 Å). Transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images were recorded by using a JEM-2100F STEM/EDS microscope. The Fourier transform infrared (FTIR) instrument used was a Nicolet 60-SXB spectrometer. X-ray photoelectron spectroscopy (XPS) measurements were conducted using a VG MultiLab 2000 instrument. CHNS element analysis was conducted on a Vario EL cube Elementar. Electrochemical measurements The sodium storage behaviours were characterized by assembly of CR2016 coin cells. The working electrodes were prepared by mixing the as-prepared products, carboxyl methyl cellulose (CMC) and acetylene black at a weight ratio of 80: 10: 10. The slurry was casted onto aluminium (Al) foil and completely dried in a vacuum oven at 70 °C overnight. The average mass loading was about 1.3 mg cm-2. Sodium metal foil was utilized as the counter electrode and reference electrode; a glass fibre membrane was used as the separator. The electrolyte was a solution of 1 M trifluomethanesulfonate (NaCF3SO3) in diethyleneglycol dimethylether (DEGDME). Galvanostatic discharge−charge tests were conducted on a multichannel battery testing system (LAND CT2001A), at a potential range of 0.01−3.0 V vs. Na+/Na. Cyclic voltammetry (CV) was tested with an Autolab CHI760E electrochemical workstation. Electrochemical impedance spectroscopy (EIS) measurements were performed on an Autolab PGSTAT 302N electrochemical workstation over the 6 ACS Paragon Plus Environment

Page 7 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

frequency range of 100 kHz–0.01 Hz. All above measurements were performed on at room temperature. RESULTS AND DISCUSSION Microstructure and morphology characterization The schematic illustration for the construction of various nanostructures is briefly depicted in Fig. 1. Firstly, Bi2O3 sheets are synthesized through a facile water-bath process at room temperature. For the synthesis of BS-BO heterostructure, TAA is utilized as the sulfurization agent and Bi2O3 sheet is used as the precursor. The BS-BO heterostructured sheets can be achieved by using an appropriate amount of sulfurization agent and suitable sulfurization time. When an excess amount of sulfurization agent is employed, the Bi2O3 sheets transform into Bi2S3 sheets completely.

Figure 1 Schematic illustration of the synthesis process of Bi2O3 sheets, Bi2S3-Bi2O3 heterostructured sheets and Bi2S3 sheets.

Fig. 2a and Fig. S1 depict the XRD patterns of Bi2O3, BS-BO heterostructure, and Bi2S3. As shown in Fig. 1a, all the diffraction peaks of Bi2O3 precursor can be indexed to the cubic phase Bi2O3 (JCPDS 052-1007). The peaks exhibit broad full width at half maximum (FWHM) and weak intensity, implying the low-crystalline 7 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 25

feature of Bi2O3. This result agrees well with previous literature that highly-crystallized cubic phase Bi2O3 is difficult to be achieved at room temperature without further annealing.28 After moderate sulfurization, the diffraction peaks for Bi2O3 still remain; meanwhile, new peaks belonging to the orthorhombic phase Bi2S3 (JCPDS 17-0320) appear. The XRD pattern of Bi2S3 is presented in Fig. S1, confirming the successful synthesis of pure orthorhombic phase bismuth sulfide (JCPDS 17-0320). No obvious peaks from Bi2O3 can be observed, suggesting that the Bi2O3 is completely converted into Bi2S3 with an enough amount of sulfurization agent.

Figure 2 (a) XRD patterns of Bi2O3 sheets and BS-BO heterostructures. (b) FTIR spectra of Bi2O3 sheets and BS-BO heterostructures. (c) XPS survey spectra of Bi2O3 sheets and BS-BO heterostructures. (d) High-resolution Bi 4f XPS spectra of Bi2O3 sheets and BS-BO heterostructures.

8 ACS Paragon Plus Environment

Page 9 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

The FTIR spectra of Bi2O3 precursor and BS-BO heterostructure are presented in Fig. 2b. The bands detected at 533, 711 and 957 cm-1 can be observed in both Bi2O3 and BS-BO. For BS-BO, the emerging new band at 619 cm-1 can be assigned to Bi–S caused by the partial transformation to Bi2S3.35, 40 It worth mentioning that the bands located at 1384 and 1626 cm-1 are attributed to C–OH and C–N stretching vibrations, respectively. These bands indicate the existence of CTAB species and H2O absorbed on the surface of the final products.41 XPS survey spectra of Bi2O3 and BS-BO heterostructure are provided in Fig. 2c. Signals from Bi, O, and C elements can be observed in the XPS spectrum of Bi2O3. After sulfurization, the S 2s peak at 225.1 eV can be detected in BS-BO, confirming the successful incorporation of S. Fig. 2d shows the high-resolution XPS spectra of Bi2O3 and BS-BO heterostructure. For Bi2O3, the two strong peaks located at 164.1 and 158.3 eV can be ascribed to Bi 4f7/2 and Bi 4f5/2, respectively. For BS-BO, the peaks with binding energy of 162.8 and 157.6 eV are correlated with Bi 4f7/2 and Bi 4f5/2 in Bi2S3, respectively. Moreover, the Bi 4f7/2 and Bi 4f5/2 peaks for Bi2O3 shift towards higher energies by approximately 0.5 eV, which is due to the coupling effect resulted from the strong chemical bonding between Bi2O3 and Bi2S3.42 The peak centred at 160.4 eV in BS-BO is ascribed to S 2p transition,43 which further confirms the formation of BS-BO heterostructures. In addition, the high-resolution C 1s (Fig. S2a) and O 1s (Fig. S2b) XPS spectra of BS-BO indicate the presence of absorbed species, such as CTAB and H2O, which are in consistent with the FTIR results. CHNS elemental analysis results are shown in Table S1; the weight percentage of sulfur in 9 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 25

BS-BO is determined to be 14.83 %. The above results reveal that the orthorhombic Bi2S3 and cubic Bi2O3 coexist in the BS-BO heterostructured composite.

Figure 3 Morphology characterizations of BS-BO heterostructures. SEM images (a, b), TEM images (c-d) and HRTEM images (e), STEM images (f, g) and corresponding elemental mapping images (from the red rectangle area in Fig. 3f) of Bi (h), S (i), O (j) of BS-BO heterostructured sheets.

The morphologies and microstructures of the BS-BO, Bi2O3 and Bi2S3 were characterized by FESEM and TEM. The Bi2O3 displays a uniform sheet-like structure with ultrathin thickness, and quite smooth surface (Fig. S3). The Bi2O3 sheets are 10 ACS Paragon Plus Environment

Page 11 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

wrinkled rather than flat, which avoids their stacking (Fig. S3a). The thickness of the Bi2O3 sheets is less than 20 nm (Fig. S3b). TAA was utilized as the sulfurization agent to convert the Bi2O3 into Bi2S3.44 The amount of sulfurization agent plays a vital role in the synthesis. If an excess amount of TAA was used, micrometer-sized Bi2S3 sheets can be obtained (Fig. S4a). The Bi2S3 sheets show a coarse surface (Fig. S4b). With the use of an appropriate amount of TAA, BS-BO heterostructured composite can be achieved (Fig. 3). The as-synthesized BS-BO well inherits the sheet-like morphology of the Bi2O3 precursor; however, it shows a coarser surface with less wrinkles (Fig. 3a). The diameter and thickness of the BS-BO nanosheets are determined to be around 600 nm and tens of nanometer, respectively (Fig. 3b). When further examined under TEM, the wrinkles can be clearly observed on the surface of BS-BO (indicated by arrows in Fig. 3c). The heterostructure feature can be clearly recognized by the contrast in the TEM image (Fig. 3d), in which the Bi2S3 shows darker mass-thickness contrast (highlighted with red circle) while the Bi2O3 shows brighter contrast. A typical HRTEM image of BS-BO is provided in Fig. 3e. The atomic spacing of 0.28 nm and 0.36 nm correspond to the (221) and (130) planes of orthorhombic Bi2S3, respectively. The lattice fringe of 0.32 nm is correlated with the (111) planes of cubic Bi2O3. The elemental mapping images under TEM (Fig. 3g-j) confirm the uniform distribution of Bi, S, and O elements within BS-BO nanosheets. The above results indicate sheet-like BS-BO heterostructures have been successfully constructed.

11 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 25

Sodium storage performance

Figure 4 Electrochemical performances of BS-BO, Bi2S3 and Bi2O3 electrodes for sodium storage. (a) The typical charge/discharge profiles of BS-BO at the current density of 100 mA g-1 for the initial three cycles. (b) Cycling performance of BS-BO at 100 mA g-1 for 20 cycles. (c) Cycling performance of BS-BO, Bi2S3 and Bi2O3 sheets at the current density of 200 mA g-1. (d) The rate performances of BS-BO, Bi2S3 and Bi2O3 sheets at different current densities.

The sodium storage behaviours of BS-BO, Bi2S3, and Bi2O3 sheets were examined in CR2016 coin cells. Cyclic voltammetry (CV) was first employed to investigate electrochemical performance of BS-BO (Fig. S5). During the first cathodic scan, a broad reduction peak located at 0.51 V can be observed. This peak splits into two peaks at around 0.31 V and 0.68 V in the subsequent scans, suggesting the evolution might be associated with irreversible activation. The first oxidation process 12 ACS Paragon Plus Environment

Page 13 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

of BS-BO is featured by two peaks at around 0.64 V and 0.78 V, which are different from the typical oxidation peaks of Bi2O3 (Fig. S6) and Bi2S3 anodes (Fig. S7). Due to the homogeneous distribution at nanoscale and strong chemical interactions between Bi2S3 and Bi2O3, the electrochemical redox process of BS-BO cannot be directly recognized as the simple overlap of two counterparts. As CV cycles proceed, the profiles of BS-BO change slightly with less pronounced current responses. The first three galvanostatic discharge-charge profiles of BS-BO, Bi2O3, and Bi2S3 at a current density of 100 mA g-1 are shown in Fig. 4a, Fig. S8 and Fig. S9, respectively. The first discharge and charge capacities for BS-BO are 821 and 571 mA h g-1, corresponding to a relatively low initial Coulombic efficiency of 69.5%. The irreversible capacity loss might be caused by the generation of SEI films. In details, the initial discharge plateau of BS-BO at around 1.55 V representing activation process is almost lost in the subsequent cycles. Two apparent discharge plateaus are located at 0.65 V and 0.45 V, agreeing well with the CV profiles. These plateaus are associated with the formation of metallic Bi and Na3Bi alloy, respectively. The discharge plateaus are consistent with the ex-situ studies on bismuth-based anode materials.29 The main charge plateau is located at 0.61 V, which is ascribed to the dealloying process from Na3Bi to metallic Bi and the oxidation of Bi to Bi3+.25 The cycling performance and corresponding Coulombic efficiency of BS-BO at a current density of 100 mA g-1 are shown in Fig. 4b. A large capacity loss occurs in the first cycle. The second discharge capacity can reach 630 mA h g-1, which is slightly higher than the theoretic capacity of Bi2S3 (625 mA h g-1) and approaches the theoretical 13 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 25

capacity of Bi2O3 (690 mA h g-1). After 20 cycles, the BS-BO is capable to deliver a discharge capacity of 477 mA h g-1, corresponding to a capacity retention of 76% (against the second discharge capacity). Additionally, the Coulombic efficiency increases significantly during the initial six cycles and stabilizes at around 98% in the subsequent processes. For comparison, the cycling performances of BS-BO, Bi2S3 and Bi2O3 at a current density of 200 mA g-1 are provided in Fig. 4c. The capacity of Bi2O3 decays fast from the start of cycling. After 50 cycles, the capacity of Bi2O3 is merely around 109 mA h g-1. A similar severe capacity decay is also observed in Bi2S3. With respect to BS-BO, the capacity decay is mild. The capacity retentions for BS-BO, Bi2S3, and Bi2O3 sheets (against second discharge capacity) are 43%, 11% and 2%, respectively. The great enhancement in capacity retention confirms the advantages of the novel BS-BO heterostructure. To further demonstrate the good sodium storage performance of BS-BO, the rate performances of the three samples were measured with the applied current densities increasing from 50 mA g-1 to 2000 mA g-1 and then reducing back to 100 mA g-1 (Fig. 4d). As can be seen clearly, the BS-BO exhibits a remarkably improved rate performance at a variety of current densities. The reversible discharge capacities are approximately 620, 574, 518, 460, 404 and 314 mA h g-1 at 50, 100, 200, 500, 1000 and 2000 mA g-1, respectively. Besides, when the current density reduces to 100 mA g-1, an average capacity of 468 mA h g-1 can be restored for the BS-BO. In contrast, both Bi2S3 and Bi2O3 sheets suffer from severe capacity loss at high rate and exhibit 14 ACS Paragon Plus Environment

Page 15 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

poor capacity recovery ability. These results demonstrate that the constructed BS-BO heterostructure is beneficial to the rate capability when compared with the Bi2S3 or Bi2O3 counterparts. Electrochemical impedance spectroscopy (EIS) measurements were performed to investigate the kinetic characteristic and charge transfer resistance of various electrodes. All typical Nyquist plots provided in Fig. S10 exhibit a slope line in the low frequency region and a semicircle in the high– medium frequency region. It can be clearly observed the semicircle diameter for BS-BO is much smaller than those for Bi2S3 and Bi2O3 sheets. After simulation, the charge transfer resistance (Rct) values for the BS-BO, Bi2S3, and Bi2O3 sheets are calculated to be 17.8, 42.6 and 209.2 Ω, respectively. The EIS results confirm the enhanced charge transport in BS-BO heterostructures.

Figure 5 Schematic illustration of the accelerated charge transfer resulted from induced electric field mechanism in BS-BO heterostructures.

15 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 25

The improved sodium storage performance originates from the novel design of BS-BO heterostructures with three distinctive merits. Firstly, a built-in electric field within heterostructure can accelerate the charge transfer. A possible mechanism of the evolution of built-in electric field in BS-BO can be illustrated in Fig. 5. The Bi2O3 is a

n-type semiconductor featured with a relative wide band gap of 2.8 eV,13, 14 while Bi2S3 is a typical p-type semiconductor with a narrow band gap (1.3 eV).28 In the BS-BO heterostructure, Bi2S3-Bi2O3 p-n heterojunctions formed at the interface can induce a built-in electric field with a direction from Bi2O3 (Eg ≈ 2.8 eV) to Bi2S3 (Eg ≈ 1.3 eV).45 It can be anticipated that during the discharge process, this electric field can facilitate the immigration of Na+ ions. On the other hand, Bi2O3 is converted into NaxBi and Na2O after full sodiation. Meanwhile, Bi2S3 is sodiated into NaxBi and Na2S.37 As evidenced in literatures, more Na+ ions might be released at sulfide area (Bi2S3-domained area) after full de-sodiation on account of its higher reversibility relative to some oxides (Bi2O3-domained area).37, 46 Consequently, when subject to charge process, an electric field could be formed with a direction from Bi2S3 to Bi2O3 owing to the potential difference, which will promote the transfer of Na+ ions. Therefore within the BS-BO heterostructure, the induced built-in electric field would provide a driving force for Na+ ions and remarkably improve the charge transfer kinetics, resulting in good rate capability. Secondly, the strong chemical interaction between Bi2O3 and Bi2S3 would be helpful to the elastic effects of the nanostructures upon sodiation/de-sodiation.47 Last but not the least, the thin nanosheets design can effectively decrease the diffusion pathway of Na+ ions, which further promotes the 16 ACS Paragon Plus Environment

Page 17 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

electrochemical kinetics. In a word, the synergetic effect of enhanced charge transfer kinetic, strong chemical interaction and nanoscale size design makes BS-BO nanosheets as a promising material for high-performance sodium-ion battery anode. CONCLUSION In summary, Bi2S3-Bi2O3 heterostructured nanosheets have been constructed via an easy water-bath approach. The present method has been demonstrated to be a TAA-directed surfactant-assisted process. When applied in sodium storage, the BS-BO heterostructure displays a promising performance: it delivers a reversible discharge capacity of 630 mA h g-1 at 100 mA g-1 with relative good stability. Moreover, the BS-BO heterostructure exhibits improved rate capability compared to Bi2S3 and Bi2O3 counterparts, which can be originated from the synergetic effect of enhanced charge transfer in the heterostructures, strong chemical interaction and nanoscale size design. This work suggests the construction of novel heterostructures can be a promising approach towards for high-performance sodium-ion battery and related energy storage applications. ASSOCIATED CONTENT

Supplementary Information. Additional Figure S1-S10 and Table S1. XRD patterns Bi2S3 sheets (Figure S1); XPS studies of C, O elements in BS-BO (Figure S2); SEM images of Bi2O3 sheets (Figure S3) and Bi2S3 sheets (Figure S4); CHNS test results of BS-BO (Table S1); CV profiles of various electrodes for the initial three cycles (Figure S5, S6 and S7); the typical charge/discharge curves of Bi2S3 sheets (Figure S8)

17 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 25

and Bi2O3 sheets (Figure S9) for the initial three cycles; Nyquist plots of various electrodes (Figure S10). This material is available free of charge via the Internet at http://pubs.acs.org. AUTHOR INFORMATION Corresponding Author *E-mail: [email protected]

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

Notes The authors declare no competing financial interest. ACKNOWLEDGMENT This work was supported by the National Key Research and Development Program of China (2016YFA0202603), the National Basic Research Program of China (2013CB934103), the Programme of Introducing Talents of Discipline to Universities (B17034), the National Natural Science Foundation of China (51521001, 21673171, 51502226), the National Natural Science Fund for Distinguished Young Scholars (51425204), and the Fundamental Research Funds for the Central Universities (WUT: 2016-JL-004, 2016III001, 2016III002, 2017III009, 2017III008), Prof. Liqiang Mai gratefully acknowledged financial support from China Scholarship Council (No. 201606955096). 18 ACS Paragon Plus Environment

Page 19 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

REFERENCES 1.Armand, M.; Tarascon, J. M. Building better batteries. Nature 2008, 451 (7179), 652−657. 2.Goodenough, J. B.; Park, K. S. The Li-ion rechargeable battery: a perspective. J.

Am. Chem. Soc. 2013, 135 (4), 1167−1176. 3. Mai, L. Q.; Tian, X. C.; Xu, X.; Chang, L.; Xu, L. Nanowire electrodes for electrochemical energy storage devices. Chem. Rev. 2014, 114 (23), 11828−11862. 4. Kim, S. W.; Seo, D. H.; Ma, X.; Ceder, G.; Kang, K. Electrode materials for rechargeable sodium-ion batteries: potential alternatives to current lithium-ion batteries. Adv. Energy Mater. 2012, 2 (7), 710−721. 5. Pan, H.; Hu, Y. S.; Chen, L. Room-temperature stationary sodium-ion batteries for large-scale electric energy storage. Energy Environ. Sci. 2013, 6 (8), 2338−2360. 6. Cao, Y. L.; Xiao, L.; Sushko, M. L.; Wang, W.; Schwenzer, B.; Xiao, J.; Nie, Z.; Saraf, L. V.; Yang, Z.; Liu, J. Sodium ion insertion in hollow carbon nanowires for battery applications. Nano Lett. 2012, 12 (7), 3783−3787. 7. Kundu, D.; Talaie, E.; Duffort, V.; Nazar, L. F. The emerging chemistry of sodium ion batteries for electrochemical energy storage. Angew. Chem. Int. Ed. 2015, 54 (11), 3431−3448. 8. Luo, W.; Zhang, P. F.; Wang, X. P.; Li, Q. D.; Dong, Y. F.; Hua, J. C.; Zhou, L.; Mai, L. Q. Antimony nanoparticles anchored in three-dimensional carbon network as promising sodium-ion battery anode. J. Power Sources 2016, 304, 340−345. 9. Li, Q. D.; Wei, Q. L.; Zuo, W.; Huang, L.; Luo, W.; An, Q. Y.; Pelenovich, V. O.; 19 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 25

Mai, L. Q.; Zhang, Q. J. Greigite Fe3S4 as a new anode material for high-performance sodium-ion batteries Chem. Sci. 2017, 8 (1), 160−164. 10. Zhang, K.; Hu, Z.; Liu, X.; Tao, Z.; Chen, J. FeSe2 microspheres as a high-performance anode material for Na-ion batteries. Adv. Mater. 2015, 27 (21), 3305−3309. 11. Denis, Y.; Prikhodchenko, P. V.; Mason, C. W.; Batabyal, S. K.; Gun, J.; Sladkevich, S.; Medvedev, A. G.; Lev, O. High-capacity antimony sulphide nanoparticle-decorated graphene composite as anode for sodium-ion batteries. Nature

Commun. 2013, 4, 2922. 12. Luo, W.; Calas, A.; Tang, C. J.; Li, F.; Zhou, L.; Mai, L. Q. Ultralong Sb2Se3 nanowire based free-standing membrane anode for lithium/sodium ion batteries. ACS

Appl. Mater. Interfaces 2016, 8, 35219−35226. 13. Biswas, K.; Zhao, L. D.; Kanatzidis, M. G. Tellurium-free thermoelectric: the anisotropic n-type semiconductor Bi2S3. Adv. Energy Mater. 2012, 2 (6), 634−638. 14. Ge, Z. H.; Zhang, B.P.; Shang, P. P.; Li, J. F. Control of anisotropic electrical transport property of Bi2S3 thermoelectric polycrystals. J. Mater. Chem. 2011, 21 (25), 9194−9200. 15. Pineda, E.; Nicho, M. E.; Nair, P.; Hu, H. Optoelectronic properties of chemically deposited Bi2S3 thin films and the photovoltaic performance of Bi2S3/P3OT solar cells

Solar Energy 2012, 86 (4), 1017−1022. 16. Yu, H.; Huang, J.; Zhuang, H.; Zhao, Q.; Zhong, X. Nanostructure and charge transfer

in

Bi2S3-TiO2

heterostructures.

Nanotechnology,

2014,

25

(21), 20

ACS Paragon Plus Environment

Page 21 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

215702−215710. 17. Brahimi, R.; Bessekhouad, Y.; Bouguelia, A.; Trari, M. Visible light induced hydrogen evolution over the heterosystem Bi2S3/TiO2. Catal. Today, 2007, 122 (1), 62−65. 18. Kim, J.; Kang, M. High photocatalytic hydrogen production over the band gap-tuned urchin-like Bi2S3-loaded TiO2 composites system. Int. J. Hydrogen Energy, 2012, 37 (10): 8249−8256. 19. Chen, G. H.; Yu, Y. Q.; Zheng, K.; Ding, T.; Wang, W. L.; Jiang, Y.; Yang, Q. Fabrication of ultrathin Bi2S3 nanosheets for high-performance, flexible, visible-NIR photodetectors. Small 2015, 11 (24), 2848−2855. 20. Konstantatos, G.; Levina, L.; Tang, J.; Sargent, E. H. Sensitive solution-processed Bi2S3 nanocrystalline photodetectors. Nano Lett. 2008, 8 (11), 4002−4006. 21. Tahir, A. A.; Ehsan, M. A.; Mazhar, M.; Wijayantha, K. U.; Zeller, M.; Hunter, A. Photoelectrochemical and photoresponsive properties of Bi2S3 nanotube and nanoparticle thin films. Chem. Mater. 2010, 22 (17), 5084−5092. 22. Ni, J. F.; Zhao, Y.; Liu, T. T.; Zheng, H. H.; Gao, L. J.; Yan, C. L.; Li, L. Strongly coupled Bi2S3@ CNT hybrids for robust lithium storage. Adv. Energy Mater. 2014, 4 (16) 1400798. 23. Zhao, Y.; Gao, D. L.; Ni, J. F.; Gao, L. J.; Yang, J.; Li, Y. One-pot facile fabrication of carbon-coated Bi2S3 nanomeshes with efficient Li-storage capability.

Nano Res. 2014, 7 (5), 765−773. 24. Li, W. J.; Han, C.; Chou, S. L.; Wang, J. Z.; Li, Z.; Kang, Y. M.; Liu, H. K.; Dou, 21 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 25

S. X. Graphite-nanoplate-coated Bi2S3 composite with high-volume energy density and excellent cycle life for room-temperature sodium-sulfide batteries. Chem. –A Eur.

J. 2016, 22 (2), 590−597. 25. Yang, W. L.; Wang, H.; Liu, T. T.; Gao, L. J. A Bi2S3@CNT nanocomposite as anode material for sodium ion batteries. Mater. Lett. 2016, 167, 102−105. 26. Sun, W. P.; Rui, X. H.; Zhang, D.; Jiang, Y. Z.; Sun, Z. Q.; Liu, H. K.; Dou, S. X. Bismuth sulfide: A high-capacity anode for sodium-ion batteries. J. Power Sources 2016, 309, 135−140. 27. Zhou, L.; Wang, W. Z.; Xu, H. L.; Sun, S. M.; Shang, M. Bi2O3 hierarchical nanostructures: controllable synthesis, growth mechanism, and their application in photocatalysis. Chem. –A Eur. J. 2009, 15 (7), 1776−1782. 28. Deng, Z. ; Liu, T. ; Chen, T. Jiang, J. ; Yang, W.; Guo, J. ; Zhao, J. ; Wang, H. ; Gao, L. Enhanced electrochemical performances of Bi2O3/rGO nanocomposite via chemical bonding as anode materials for lithium ion batteries. ACS Appl. Mater. &

Interfaces, 2017, 9, 12469-12477. 29. Kim, M. K.; Yu, S. H.; Jin, A.; Kim, J.; Ko, I. H.; Lee, K. S.; Mun, J.; Sung, Y. E. Bismuth oxide as a high capacity anode material for sodium-ion batteries. Chem.

Commun. 2016, 52 (79), 11775−11778. 30. Nithya, C. Bi2O3@ reduced graphene oxide nanocomposite: An anode material for sodium-ion storage. ChemPlusChem 2015, 80 (6), 1000−1006. 31. Yin, H. ; Cao, M. L. ; Yu, X. X. ; Zhao, H. ; Shen, Y. ; Li, C. ; Zhu, M. Q. Self-standing Bi2O3 nanoparticles/carbon nanofiber hybrid films as a binder-free 22 ACS Paragon Plus Environment

Page 23 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

anode for flexible sodium-ion batteries. Mater. Chem. Front., 2017, 1, 1615-1621. 32. Yang, W.; Wang, H.; Liu, T.; Gao, L. A Bi2S3@ CNT nanocomposite as anode material for sodium ion batteries. Mater. Lett., 2016, 167: 102-105. 33. Gong, Y. J.; Lin, J. H.; Wang, X. L.; Shi, G.; Lei, S. D.; Lin, Z.; Zou, X. L.; Ye, G. L.; Vajtai, R.; Yakobson, B. I.; Yakobson, B. I.; Terrones, H.; Terrones, M.; Tay, B. K.; Lou, J.; Pantelides, S. T.; Liu, Z.; Zhou, W.; Ajayan, P. M. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nature Mater. 2014, 13 (12), 1135−1142. 34. Gao, X., Wu, H. B., Zheng, L., Zhong, Y.; Hu, Y.; Lou, X. W. Formation of mesoporous

heterostructured

BiVO4/Bi2S3

hollow

discoids

with

enhanced

photoactivity. Angew. Chem. Int. Ed. 2014, 53 (23), 5917−5921. 35. Liu, T. T.; Zhao, Y.; Gao, L. J.; Ni, J. F. Engineering Bi2O3-Bi2S3 heterostructure for superior lithium storage. Sci. Rep. 2015, 5, 9307. 36. Cai, F. G.; Yang, F.; Jia, Y. F.; Ke, C.; Cheng, C. H.; Zhao, Y. Bi2S3-modified TiO2 nanotube arrays: easy fabrication of heterostructure and effective enhancement of photoelectrochemical property. J. Mater. Sci. 2013, 48 (17), 6001−6007. 37. Zheng, Y.; Zhou, T. ; Zhang, C. F.; Mao, J. F.; Liu, H. K.; Guo, Z. P. Boosted charge transfer in SnS/SnO2 heterostructures: toward high rate capability for sodium-ion batteries Angew. Chem. Int. Ed. 2016, 55 (10), 3408−3413. 38. Zheng, Y. ; Zhou, T. ; Zhao, X. ; Pang, W. K.; Gao, H.; Li, S.; Zhou, Z.; Liu, H. K.; Guo, Z. P. Atomic interface engineering and electric-field effect in ultrathin Bi2MoO6 nanosheets for superior lithium ion storage. Adv. Mater. 2017, 29 (26):1700396. 23 ACS Paragon Plus Environment

ACS Applied Materials & Interfaces 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 25

39. Yan, C.; Zhu, Y.; Li, Y.; Fang, Z.; Peng, L.; Zhou, X.; Chen, G.; Yu. G. H. Local built-in electric field enabled in carbon-doped Co3O4 nanocrystals for superior lithium-ion storage. Adv. Funct. Mater. 2018, 201705951. 40. Zhang, L. X.; Li, N.; Jiu, H. F.; Zhang, Q. Solvothermal synthesis of reduced graphene oxide-Bi2S3 nanorod composites with enhanced photocatalytic activity under visible light irradiation. J. Mater. Sci. –Mater. El. 2016, 27 (3), 2748−2753. 41. Zhao, Y.; Liu, T. T.; Xia, H.; Zhang, L.; Jiang, J. X.; Shen, M.; Ni, J. F.; Gao, L. J. Branch-structured Bi2S3–CNT hybrids with improved lithium storage capability J.

Mater. Chem. A 2014, 2 (34), 13854−13858. 42. Gao, M. R.; Yu, S. H.; Yuan, J.; Zhang, W.; Antonietti, M. Poly (ionic liquid)-mediated morphogenesis of bismuth sulfide with a tunable band gap and enhanced electrocatalytic properties. Angew. Chem. Int. Ed. 2016, 128 (41), 13004−13008. 43. Tian, L.; Tan, H. Y.; Vittal, J. J. Morphology-controlled synthesis of Bi2S3 nanomaterials via single-and multiple-source approaches. Crys. Growth Des. 2007, 8 (2), 734-738. 44. Li, L. S.; Sun, N. J.; Huang, Y. Y.; Qin, Y.; Zhao, N. N.; Gao, J. N.; Li, M. X.; Zhou, H. H.; Qi, L. M. Topotactic transformation of single-crystalline precursor discs into disc-like Bi2S3 nanorod networks. Adv. Funct. Mater. 2008, 18 (8), 1194−1201. 45. Ma, D. K.; Guan, M. L.; Liu, S. S.; Zhang, Y. Q.; Zhang, C. W.; He, Y. X.; Huang, S. M. Controlled synthesis of olive-shaped Bi2S3/BiVO4 microspheres through a limited

chemical

conversion

route

and

enhanced

visible-light-responding 24

ACS Paragon Plus Environment

Page 25 of 25 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

photocatalytic activity. Dalton Transactions 2012, 41 (18), 5581−5586. 46. Cai, K.; Song, M. K.; Cairns, E. J.; Zhang, Y. Nanostructured Li2S–C composites as cathode material for high-energy lithium/sulfur batteries. Nano Lett. 2012, 12 (12), 6474−6479. 47. Ni, J. F.; Fu, S. D.; Wu, C.; Maier, J.; Yu, Y.; Li, L. Self-supported nanotube arrays of sulfur-doped TiO2 enabling ultrastable and robust sodium storage. Adv. Mater. 2016, 28, 2259–2265.

Table of Contents (TOC) graphic

25 ACS Paragon Plus Environment